Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2025 Feb 26;19(9):8895–8903. doi: 10.1021/acsnano.4c16399

Electron–Phonon Coupling and Phonon Dynamics in Single-Layer NbSe2 on Graphene: The Role of Moiré Phonons

Amjad Al Taleb , Wen Wan , Giorgio Benedek ‡,§, Miguel M Ugeda ‡,∥,, Daniel Farías †,#,∇,*
PMCID: PMC12129249  PMID: 40009749

Abstract

The interplay between substrate interactions and electron–phonon coupling in two-dimensional (2D) materials presents a significant challenge in understanding and controlling their electronic properties. Here, we present a comparative study of the structural characteristics, phonon dynamics, and electron–phonon interactions in bulk and monolayer NbSe2 on epitaxial bilayer graphene (BLG) using helium atom scattering (HAS). High-resolution helium diffraction reveals a (9 × 9)­0° superstructure within the NbSe2 monolayer, commensurate with the BLG lattice, while out-of-plane HAS diffraction spectra indicate a low-corrugated (3√3 × 3√3)­30° substructure. By monitoring the thermal attenuation of the specular peak across a temperature range of 100 to 300 K, we determined the electron–phonon coupling constant (λHAS) as 0.76 for bulk 2H-NbSe2. In contrast, the NbSe2 monolayer on graphene exhibits a reduced λHAS of 0.55, corresponding to a superconducting critical temperature (T C) of 1.56 K according to the MacMillan formula, consistent with transport measurement findings. Inelastic HAS data provide, besides a set of dispersion curves of acoustic and lower optical phonons, a soft, dispersionless branch of phonons at 1.7 meV, attributed to the interface localized defects distributed with the superstructure period, thus termed Moiré phonons. Our data show that Moiré phonons contribute significantly to the electron–phonon coupling in monolayer NbSe2. These results highlight the crucial role of the BLG in the electron–phonon coupling in monolayer NbSe2, attributed to enhanced charge transfer effects, providing valuable insights into substrate-dependent electronic interactions in 2D superconductors.

Keywords: electron−phonon coupling, phonons, NbSe2 , graphene, superconductor, Moiré structure


graphic file with name nn4c16399_0010.jpg


graphic file with name nn4c16399_0008.jpg

1. Introduction

Layered transition metal chalcogenides (TMD) are well-suited systems to study the effects of dimensionality and thickness on collective electronic phases like superconductivity and charge density waves (CDW). These are generally manifestations of electron–phonon interactions, leading to electron–electron (hole–hole) pairing and electron–hole ordering, respectively. Bulk 2H-NbSe2 is especially interesting due to the coexistence of superconducting and CDW phases: a CDW with a (3 × 3) periodicity is known to set in below 33 K, while superconductivity sets in at T c = 7.2 K. In the monolayer NbSe2 limit, different results have been reported depending on the supporting substrate. On epitaxial bilayer graphene (BLG) on SiC(0001), T C is depressed to 1.5 K whereas the critical temperature for the appearance of the CDW remains unchanged (33 K). However, when the monolayer is held on sapphire, superconductivity was observed to set in at T c = 3 K, while the CDW critical temperature increases to 145 K. These results were interpreted in terms of an increase in the electron–phonon coupling in single-layer NbSe2/sapphire. Regarding charge transfer, there is little difference between sapphire (work function ∼4.5 eV) and BLG/SiC (work function 4.30 eV), when compared to the NbSe2 work function of 5.9 eV. Thus, these different behaviors in the presence of similar charge transfer raise the question about the possible role of BLG underneath in both superconductivity and CDW and more specifically, on the strength of the electron–phonon coupling in single-layer NbSe2.

In this work, we report measurements of the electron–phonon coupling constant (mass-enhancement factor) λHAS in single-layer NbSe2/BLG/SiC­(0001) by high-resolution helium atom scattering (HAS) spectroscopy, which enabled us to extract information about the surface structure on a long-period (nm) scale and related low-energy (meV) dynamics. Our He-diffraction results show that single-layer NbSe2 on BLG develops a commensurate superstructure, where a (9 × 9) NbSe2 monolayer supercell matches with a small tensile strain a (13 × 13) BLG supercell. A comparatively intense, dispersionless low-energy (1.7 meV) phonon branch is also observed. This effect induced by the superstructure on phonon dynamics is analogous to what observed in the electronic structure, like the occurrence of ultraflat electronic bands, e.g., in twisted bilayers of WSe2 and of graphene, , ultimately responsible for the observed superconductivity in these heterostructures.

A recent quantum-theoretical approach showed how the thermal attenuation of the HAS specular peak from metal surfaces, described by the Debye–Waller (DW) exponent, can directly provide the electron–phonon coupling constant (mass-enhancement factor λ, here denoted λHAS when obtained with this method) This method to derive λHAS has been recently extended to other classes of conducting surfaces, like those of layered TMDs. , and supported graphene monolayers, and is here applied to the present NbSe2 heterostructures to elucidate the role of the low-energy phonon branch in the electron–phonon coupling. In general, HAS spectroscopy should be relevant for the characterization of twisted layered heterostructures and their functional properties in the new area of twistronics.

2. Results and Discussion

2.1. Diffraction

Monolayers of NbSe2 were grown on BLG on 6H-SiC(0001) by molecular beam epitaxy in the homemade UHV-MBE system at the DIPC in San Sebastián. Samples with a coverage of 2–3 layers of NbSe2 have also been prepared. Before being shipped to Madrid for HAS measurements, the samples were capped with a ∼10 nm film of Se (see Supporting Information for more details on sample preparation). Figure a shows the HAS angular distribution measured for single-layer NbSe2/BLG/SiC­(0001) (red curve) along the (1000) direction, and is compared to that for a 2–3 layers of NbSe2/BLG/SiC­(0001) (green curve) and for the bulk NbSe2(0001) surface (black curve), all recorded at T = 100 K. Note that these data were recorded with a HAS system that allows the detector to rotate 200° in the scattering plane while keeping the angle of incidence fixed. The spectrum for bulk NbSe2 shows numerous intense diffraction peaks, indicating a highly corrugated surface, similar to those typically found on semiconductor surfaces. A comparable diffraction spectrum is observed for the few-layer sample, while the monolayer NbSe2 exhibits a significantly different pattern. In this case, the diffraction intensities are much weaker than the specular scattering, suggesting a metallic surface for the monolayer due to significant charge transfer from the substrate. This is consistent with the comparatively large difference between the work function of bulk NbSe2 (5.9 eV) [7] and that of few-layer NbSe2/BLG (4.30 eV).

1.

1

(a) Angular distribution of HAS along the (1000) direction (ΓM) of the surfaces of bulk NbSe2(0001) (black curve), few-layer NbSe2/BLG/SiC­(0001) (green curve) and single-layer NbSe2/BLG/SiC­(0001) (red curve). The angle of incidence is 60° and the incident energy of the He beam is E i = 44 meV. (b) A factor-4 expansion of the diffraction spectrum for the monolayer NbSe2, where the BLG (01̅) diffraction peak is also visible, as an effect of the charge transfer.

The charge transfer is also responsible of a small shift of the diffraction peaks of the NbSe2 monolayer as compared to those of the bulk and few-layer samples. For the (01̅) diffraction the latter occur at θ f = 39.5°, which gives an in-plane lattice constant a N = 3.44 Å, in perfect agreement with the value measured in NbSe2 multilayer flakes, while the (01̅) peak maximum for the monolayer is shifted to 39.9°. This corresponds to a N = 3.52 Å, i.e., to a ∼ 2.3% expansion of the in-plane monolayer lattice constant. The electronic charge transfer leaves behind a positive potential in the substrate and, therefore, a component in the surface charge density distribution reflecting the substrate periodicity. Actually, the ×4 magnification of the diffraction in the monolayer (Figure b) hints to a maximum around θ f = 33°, where the BLG (01̅) diffraction peak is expected to occur for the known graphene lattice parameter of 2.46 Å. ,

The observation of the diffraction peaks of both the monolayer NbSe2 and the BLG substrate induced by a large charge transfer is confirmed by the high-resolution HAS diffraction pattern displayed in Figure . This spectrum was measured using the HAS instrument with a time-of-flight (TOF) arm and a fixed angle of 108° between the incoming and outgoing beams. In this setup, angular distributions are measured by rotating the crystal around an axis perpendicular to a plane defined by the incoming beam and the normal to the sample surface. Throughout this paper we will use the term rocking angle (φ) which means the scattering angle relative to the specular position. The (01̅) BLG diffraction peak at φ = −22.1° is about one-half in size as that of single-layer NbSe2 at φ = −15.2° and comparable in sharpness, and corresponds to the in-plane lattice constants a BLG = 2.47 ± 0.02 Å and a N = 3.55 ± 0.02 Å. This value of a N confirms the in-plane expansion (here 3.2 ± 0.6%) of the NbSe2 monolayer with respect to the bulk induced by the charge transfer.

2.

2

High-resolution HAS angular distribution measured as a function of the rocking angle (top-left inset) along the (1000) direction (ΓM) of single-layer NbSe2/BLG/SiC­(0001) at T = 100 K, with a fixed scattering angle of 108° and incident energy of 43 meV. The upper part of the figure shows the corresponding diffraction spots in the reciprocal space of single-layer NbSe2 (black dots) and those near the specular peak of the single-layer NbSe2/BLG superstructures indicated in the top-right inset and discussed in the text.

Since the diffraction components of both the single-layer NbSe2 and BLG are observed in Figure , the small features at φ ≅ −5, −3.3, −1.7 and 5° near the specular peak and discernible above the noise (especially in the filtered data, red line), are assigned to a commensurate superstructure with a unit cell aligned with those of the two component lattices. The lattice constant ratio a N/a BLG = 1.44 is very close to 13/9 = 1.4̅, suggesting that the hexagonal superlattice unit cell is formed by the coincidence of a (9 × 9)­0° cell of the NbSe2 monolayer with a (13 × 13)­0° cell of BLG. This is illustrated by Figure a, which shows the interface Se ions (green atoms) positioned over the graphene (gray atoms), with the superlattice unit cell edges marked in red. The corresponding theoretical diffraction spots around the origin (specular peak) in the reciprocal plane are shown in the upper part of Figure (red and green dots), with some spots closely corresponding to the small features described above.

3.

3

(a) Structure of the single-layer NbSe2/BLG interface showing the Se ions (green full circles) on graphene (gray C atoms). The red lines show the unit cell of the single-layer NbSe2 (9 × 9)­0° supercell matching the (13 × 13)­0° supercell of BLG. (b) The reduced (3√3 × 3√3)­30° unit cell of single-layer NbSe2 (green contour line) in case of equivalence between the BLG high-symmetry hexagonal sites (left and right corners of the reduced cell) and the trigonal symmetry sites (top and bottom corners), approximately produced by the addition of the second BLG graphene layer (not shown).

The presence of a second graphene layer underneath causes the Se ions at the hexagonal symmetry sites (e.g., the corners of the (9 × 9)­0° unit cell in Figure a) to be almost equivalent, in terms of adsorption energy, to those at the trigonal symmetry sites (e.g., the top and bottom atoms of the reduced (3√3 × 3√3)­30° unit cell, as shown in Figure b). If these sites were strictly equivalent, the actual (3√3 × 3√3)­30° supercell would be three times smaller, and only the diffraction peaks corresponding to the green spots in Figure would appear. However, even with approximate equivalence, the green spotsespecially the first six, which form a 30° rotated hexagon in Figure are reinforced.

Out-of-plane measurements made using the HAS confirm this effect. The HAS intensity, mapped as a function of the final angle θ f and out-of-plane azimuth angle ϕ f for a fixed incident energy E i = 56 meV and incident angle θ i = 60° (see Figure b; kinematics explained in Figure a), shows maxima at the expected positions for the (3√3 × 3√3)­30° supercell (green hexagons). Note that the hexagonal pattern is slightly rotated clockwise relative to the calculated maxima positions, due to a minor azimuthal misalignment of the NbSe2 lattice with respect to the sagittal plane ϕ i = 0° (see Figure SI-1 for details). However, this small instrumental misalignment does not undermine the strong evidence supporting the (3√3 × 3√3)­30° supercell as the approximate commensurate structure for single-layer NbSe2/BLG/SiC­(0001).

4.

4

(a) Geometry and kinematics of out-of-plane HAS diffraction relating the components of the final parallel wavevector K f and parallel wavevector transfer ΔK to the incident wavevector k i (or incident energy through k i –1] = 1.383√E i [meV] for 4He atoms) and scattering geometry. (b) The out-of-plane HAS diffraction map of single-layer NbSe2/BLG/SiC­(0001) at T = 40 K, measured as a function of the final and azimuthal angles defined in (a) at a fixed incident angle of 60° and energy of 52 meV. The 30°-rotated hexagonal pattern reflects the theoretical positions of the diffraction peaks (green hexagonal dots) for the (3√3 × 3√3)­30° superlattice.

It is worth mentioning that essentially the same diffraction spectra for the NbSe2 monolayer, including the out-of-plane map demonstrating the (3√3 × 3√3)­30° superlattice, have been measured at T = 40 K (see Figure SI-1), which indicates that there is no evidence for a (3 × 3) CDW pattern above 40 K. This is in agreement with previous STM results on the same system, which showed that a CDW sets in below 35 K, and in clear contrast with previous Raman measurements using samples exposed to ambient conditions where a critical temperature of 145 K has been reported.

2.2. Debye–Waller Exponent

Figure shows the temperature dependence of the Debye–Waller exponent, as derived from the ratio of He specular intensity at the surface temperature T, I(T) ≡ I, to that at the lowest measured temperature T 0, I(T 0) ≡ I 0, for single-layer NbSe2 compared to the data for bulk NbSe2 and the few-layer NbSe2. The data for the NbSe2 overlayers are reported for two different He beam incident energies and angles.

5.

5

Thermal attenuation of He specular intensity measured along the ΓM azimuth from bulk NbSe2 (top), few-layer NbSe2/BLG/SiC­(0001) (middle), and for single-layer NbSe2/BLG/SiC­(0001) (bottom). Note the different ordinate scales for the three sets of data. The corresponding incident energy and incident angles are given for each set of data. The straight lines correspond to the best linear fits according to eq .

While the exponential attenuation of the specular intensity is about the same for the semi-infinite crystal and for the 2–3 layers, a much steeper attenuation is observed for the NbSe2 monolayer on top of the graphene bilayer. This is mainly due to the graphene substrate, which provides several intervalley channels, as discussed in previous HAS studies on the electron–phonon interaction of graphene on various substrates. In the next subsection, the corresponding values of the electron–phonon coupling constant are derived for the present set of data, based on the information from phonon dispersion curves derived from HAS TOF data. While HAS data for bulk NbSe2 phonons are already available and assumed to be approximately valid for multilayers, the HAS phonon data for the monolayer NbSe2 are presented in the following subsection.

2.3. Phonons

Two of the several sets of HAS TOF spectra of 1 ML-NbSe2/BLG/SiC­(0001), measured at a surface temperature of 100 K along the ΓM direction, are shown in Figure a,b as functions of the energy transfer ΔE i . The inelastic peaks in the TOF spectra, like those marked by small triangles in Figure a,b for phonon creation (ΔE i < 0) and annihilation (ΔE i > 0) processes, provide the phonon energy ℏω = |ΔE i | as a function of the parallel wavevector Q = |ΔK| (full black circles in Figure c). A possible set of dispersion curves is represented by the eye guidelines (gray full lines) drawn in Figure c. They are compared to the data for the 2H-NbSe2(0001) surface (red open symbols) and to the bulk NbSe2 dispersion curves (blue broken lines).

6.

6

(a,b) Selected time-of-flight spectra measured with HAS at constant scattering angle θ i + θ f = 108°, incident energy E i = 26 meV, along the ΓM direction of single-layer NbSe2/BLG at 100 K; each spectrum is labeled by the rocking angle φ of the surface plane, so that θ i = 54° + φ and θ f = 54° - φ. (c) The corresponding phonon dispersion curves (full black circles, with full gray lines as eye guidelines) are compared to those measured with HAS for the semi-infinite NbSe2(0001) surface at room temperature (red open symbol), and to the bulk NbSe2 dispersion curves (blue broken lines). Besides the surface branches S1, S3, S4 v and S4 h, a flat soft branch S0 is observed in single-layer NbSe2/BLG at 1.7 meV, attributed to interface localized (Moiré) modes with vertical (z) polarization. The horizontal (x-polarized) counterpart is associated with the S x branch at 8 meV, which converts via avoided crossing to the S3 branch at about ΓM/2. The displacement patterns of the atomic planes for the different branches are schematically shown in (d), with the BLG illustrated by blue segments, the NbSe2 three atomic layers by gray and black segments, and the SiC(0001) substrate, assumed to be rigid, by gray boxes.

In the long-wave limit the surface phonon branches in the acoustic region of 1 ML-NbSe2/BLG labeled as S1, S3, S4 v and S4 h correspond to the Rayleigh wave, the longitudinal resonance, the shear-vertical and longitudinal oscillations of the NbSe2 monolayer against the BLG (Figure d). In 2H-NbSe2(0001) the surface unit cell includes two NbSe2 monolayers, while in our system the BLG plays, to some extent, the role of the second monolayer. This may alone explain the softening of the S1 branch and the stiffening of S3, both in the second half of ΓM, with respect to those of bulk 2H-NbSe2(0001) (red open symbols), although the stiffening of S3 may also be related to the avoided crossing with S x . Note that a possible Kohn anomaly occurs in the S1 branch near 0.6 Å–1, more pronounced than that found with HAS at room temperature in bulk NbSe2(0001), and showing a shift to a smaller Q with respect to that reported with neutron scattering in bulk at 2/3ΓM (blue broken line). This is probably related to the charge transfer and the consequent shift of the Fermi wavevector. A similar shift of the anomaly has been observed with HAS in 2H-TaSe2(0001) for the Rayleigh mode with respect to the anomaly in the bulk acoustic mode.

The low-energy flat phonon branch S0, displaying a strong intensity at small Q (Figure a.b) may be associated with soft shear-vertical vibrations localized at a superlattice periodic array of atoms, for example the interface Se ions which accommodate at the high symmetry hexagonal (and possibly trigonal) sites of graphene (corner atoms in Figure b) with strongly weakened local shear force constants. Note that this branch is reminiscent of the low energy dispersionless modes observed on Gr/Ir(111) and Gr/Ru(0001), where graphene builds a Moiré pattern. , Due to this analogy and the present interpretation of the flat branch as closely associated with the interface superstructure, in the following discussion we shall term these localized excitations as Moiré phonons. On the other hand, such localized shear-vertical vibrations should have their longitudinal companion, which is here associated with the S x additional branch, also localized above the maximum of the acoustic band in the region around the zone center. Clearly, only a surface dynamics first-principle calculation for this system would allow for a reliable assignment of the observed additional Moiré branches S0 and S x .

2.4. Electron–Phonon Interaction

The HAS data for ln­(I/I 0), shown in Figure for the surfaces studied, provide the HAS Debye–Waller (DW) exponent −2W(T), up to a constant that depends on how I 0 is defined. Therefore, the slope of ln­(I/I 0) as a function of temperature directly represents the slope of −2W(T), regardless of the specific definition of I 0. This slope can be used to determine the electron–phonon coupling constant, λHAS (also known as the mass-enhancement factor), for bulk NbSe2 and the few layers. This approach is based on the method valid in the high-temperature limit.

For single-layer NbSe2/BLG/SiC, at least two different contributions to λHAS are expected from phonons from the NbSe2 monolayer and Moiré phonons from the interface. The high-energy optical phonons of the BLG can be neglected, however, due to their small population in the present temperature range, and their depth beneath the first surface atomic layer. In this approximation, the expression of the DW exponent receiving contributions from two distinct phonon spectral regions can be written as in ref .

2W(ki,T)=ackiz2πϕnsλHAS{Aω0coth(ω02kBT)+(1A)ω1coth(ω12kBT)} 1

where ω0 = 1.7 meV and ω1 = 13.5 meV are the frequency of Moiré phonons and the surface Debye frequency of bulk NbSe2 (taken as the bulk Debye frequency divided by √2), respectively, and the coefficient 0 ≤ A ≤ 1 weighs the two contributions. Although the Moiré phonons fall in the spectral region of acoustic phonons, their flat, optical-like dispersion, giving a sharp peak at ω0 in the phonon density, need to be treated separately from the acoustic phonons, which have instead a linear dispersion and are collectively represented by a surface Debye frequency ω1 ≫ ω0. In the prefactor of eq r.h. member, a c is the surface unit cell area, equal to 10.276 Å2 for the semi-infinite bulk NbSe2 and for the few layers, while for single-layer NbSe2 the value a c = 10.914 Å2 is determined from the measured lattice parameter including the dilation induced by charge transfer. Moreover, k iz is the surface normal component of the incident He atom wavevector and ϕ the surface workfunction. For the semi-infinite NbSe2 the work function is ϕ = 5.9 eV. For the nML-NbSe2 film, the small reduction of the HAS diffraction intensities with respect to those of the semi-infinite NbSe2 (Figure ) indicate a similarly small decrease of the surface corrugation due to a charge transfer from the SiC(0001) substrate, whose work function is ϕ = 4.6 eV. This value is used for the nML, while the even smaller workfunction of the substrate BLG/SiC(0001), ϕ = 4.3 eV, is used for the NbSe2 monolayer, which undergoes alone a massive charge transfer and therefore a flattening of the surface. Finally, the factor n s in eq represents the number of layers involved in the electron–phonon interaction probed by the scattering He atom. It stems from the experimental observation that for ultrathin conducting films the slope of 2W as a function of T grows proportionally with the number n of layers and saturates to a constant n sat . This leads to the assumption that, as far as the DW exponent is concerned, a multilayer film may be viewed as a stack of 2D free-electron gases. In the present case, the nearly equal DW slopes in Figure for nML (n = 2–3) and the semi-infinite NbSe2 indicate that also n sat is in the range 2–3. In the following n s = 2 shall be used for the semi-infinite and nML-NbSe2, consistently with the previous analysis of other 2H-stacked transition metal dichalcogenides (TMDs). ,, The large increase of the DW slope observed for 1 ML/BLG/SiC(0001) is then attributed to the insertion of BLG. The model formulated above for the origin of Moiré phonons implies that also the graphene layer interfaced to the 1 ML NbSe2 has to contribute to the electron–phonon interaction via intra- as well as interpocket (Dirac cone) transitions. As discussed in, such a multivalley transition multiplicity contributes six units to the total n s, which is therefore n s = 7. The second graphene layer acts as a buffer layer passivating the SiC(0001) surface and is assumed to play no role.

In the following analysis of the DW slopes based on eq , for a given value of A, the mass-enhancement factor λHAS works as a free parameter so as to obtain the best fit of the −2W(T) shape (equal to that of ln­(I(T)/I 0)). In this way λHAS is obtained. The high-temperature limit is equivalent to letting ω0 and ω1 in eq tend to zero, which gives

2W(ki,T)=2ackiz2πϕnsλHASkBT 2

and therefore

λHAS=πϕ2nsackiz2ln[I(T)/I0]kBT 3

as reported in ref .

The best linear fits in the high-T limit (eq ) for all the considered samples are shown in Figure (straight lines) and give λHAS = 0.76 ± 0.06 for semi-infinite bulk NbSe2, λHAS = 0.82 ± 0.06 and 0.75 ± 0.06 for the 2–3 ML-NbSe2 data sets 43 meV/65° (k iz 2 = 14.75 Å–2) and 36 meV/60° (k iz 2 = 17.28 Å–2), respectively. For the NbSe2 monolayer on BLG/SiC the fit gives λHAS = 0.58 ± 0.04 or 0.63 ± 0.04 when measured with a 43 meV He beam at 60° incidence or with a 22 meV He beam at 52.7° incidence, respectively.

As discussed above in connection with eq , the λHAS for the two data sets of single-layer NbSe2 actually receives contributions from two distinct phonon spectral regions: Moiré phonons and the lattice dynamics of the NbSe2 layer. Figure shows three fits based on eq of the 43 meV/60° data, one with only the Moiré phonons contribution (A = 1), one with an equal share of the Moiré and the NbSe2 phonons (A = 0.5) and the third with only the NbSe2 monolayer phonons (A = 0). The two fits including NbSe2 phonons (A = 0.5, 0) yield both λHAS = 0.60 ± 0.04, which is consistent with the high-T limit for the 43 meV/60° data. All three fits adequately account quite well for the experimental slope above 150 K, where the high-T linearity is established. However, below this temperature the two fits for A ≠ 0 diverge from the one including only moiré phonons and from experiment. This suggests that the moiré phonons make a significant contribution to the electron–phonon interaction at low temperatures. In a sense, they serve as acoustic surface modes that exist in bulk NbSe2, but lose their character and localization at small ΔK in the film due to penetration into the much stiffer substrate. This specific role of the Moiré phonons, which preserve their localization also for Q → 0, is consistent with the observation that the Moiré phonons are detected by HAS in those regions of the parallel momentum transfer where they intersect the S1 and S3 branches.

7.

7

Fits of the Debye–Waller data for HAS incident energy E i = 43 meV and angle θ i = 60° using eq . The red curve represents a fit including only the Moiré phonons contribution (A = 1), the blue curve represents an equal share of the Moiré and the NbSe2 phonons (A = 0.5) and the gray curve includes only the NbSe2 ML phonons (A = 0). The few data for E i = 22 meV and θ i = 52.7°, all above 125 K, are well fitted by the high-T straight line, independently of A.

The above value of λHAS for the semi-infinite NbSe2 of 0.76 ± 0.06, compares well with the values obtained from photoemission measurements at low temperature by Valla et al. (0.85 ± 0.15) and by Rahn et al. (in the range 0.7–0.9). For the 2–3 ML the above values of λHAS = 0.82 ± 0.06 and 0.75 ± 0.06 turn out to be larger than that obtained by Lian et al. from first-principles for a Na-intercalated NbSe2‑bilayer (λ = 0.57), where the Na atoms act as electron donors, like the SiC(0001) substrate in present experiments. A more instructing comparison is for the NbSe2 monolayer on BLG/SiC, whose average value of λHAS = 0.61 ± 0.04 is somewhat smaller than the values from first-principles calculations for a NbSe2 monolayer by Zheng and Feng (λ = 0.67), by Lian et al. for two different CDW configurations (λ = 0.84 and 1.09). It is also smaller than the λ ≈ 0.75 derived by Xi et al. from electrical transport and optical measurements for a NbSe2 monolayer on sapphire (work function 4.5 eV).

2.5. Superconducting Critical Temperature

The superconducting critical temperature Tc can be derived from λHAS via the Allen–Dynes-modified McMillan formula ,

Tc=11.5KmeVωmeVexp[1.04(1+λHAS)λHASμ*0.62λHASμ*] 4

with the Coulomb pseudopotential μ* = 0.15 and ⟨ω⟩meV = 12.61 meV taken from. For the semi-infinite NbSe2HAS = 0.76 ± 0.06) it is found T c = 4.9 ± 0.6 K, which is smaller than the bulk value (7.3 K). Comparable values of T c are found from the two sets of data for few-layer NbSe2 (6.0 ± 0.7 K from λHAS = 0.82 ± 0.06 and 4.7 ± 0.6 K from λHAS = 0.75 ± 0.06) when the same values of μ* and ⟨ω⟩meV are used for the NbSe2 multilayers. Encapsulated NbSe2 multilayers with 2 ≤ n ≤ 8 have shown a uniform decrease of T c with thickness just in that range, while the T c of the monolayer encapsulated with hBN drops to 2 K.

For monolayer NbSe2 (curve labeled 43 meV/60° in Figure ), the best-fit value λHAS = 0.58 ± 0.04, obtained by including the Moiré phonons (A = 1), yields, for ⟨ω⟩meV = 12.61 meV, a smaller value of T c, equal to 1.84 ± 0.20 K. This compares well with previous experimental T C values obtained for single-layer NbSe2/BLG/SiC using STM/STS (T c = 1.5 K) and transport measurements T c = 1.5 K. Despite the possible contributions of the Moiré phonons, the graphene interfacing has apparently the effect of depressing superconductivity.

3. Conclusions

High-resolution HAS diffraction data from single-layer NbSe2 on BLG/SiC(0001) reveal a (9 × 9)­0°, approximately reducible to a (3√3 × 3√3)­30° superstructure, commensurate with the underlying BLG lattice. Inelastic HAS data provide, besides a set of dispersion curves of acoustic and lower optical phonons, a soft, dispersionless branch of phonons at 1.7 meV, attributed to the interface localized defects distributed with the superstructure period, and thus termed Moiré phonons. The electron–phonon coupling for single-layer NbSe2/BLG has been derived from the temperature dependence of the DW exponent, and compared with those of few-layer NbSe2/BLG/SiC­(0001) and bulk NbSe2. The low-temperature behavior of the DW exponent suggests an appreciable contribution of the Moiré phonons to the electron–phonon coupling, although the intercalation of the BLG between the NbSe2 monolayer and the SiC substrate is seen to attenuate the electron–phonon coupling with respect to the case of a NbSe2 monolayer directly deposited on an inert substrate.

In conclusion, it has been demonstrated that high-resolution HAS diffraction studies can be advantageously extended to long-period conducting surfaces, e.g., those exhibiting long-period superstructures, in order to obtain a detailed structural information. Moreover, the electron–phonon interaction obtained from the analysis of the HAS-DW factor as a function of temperature proves to be, within the experimental uncertainties, quite reliable, also in the case of particularly complex superconducting 2D structures. The growing interest in twisted Moiré structures, realized through the assembling of 2D layered materials, should find in HAS a valuable tool for their structural and dynamical characterization.

4. Methods

4.1. Sample Preparation

Monolayers of NbSe2 were grown on epitaxial bilayer graphene (BLG) on 6H-SiC(0001) by molecular beam epitaxy at a base pressure of ∼3 × 10–10 mbar in our homemade UHV-MBE system at the DIPC in San Sebastián (Spain). SiC wafers with resistivities ρ ∼ 120 Ω cm were used. Graphitization of the SiC surface was carried out using an automatized cycling mechanism where the sample was ramped between 700 and 1350 °C at a continuous ramping speed of ∼20 °C/s. The SiC crystal was kept for 30 s at 1350 °C for 80 cycles.

Reflective high energy diffraction (RHEED) was used to monitor the layer growth of NbSe2. During the growth, the BLG/SiC substrate was kept at 570 °C. High purity Nb (99.99%) and Se (99.999%) were evaporated using an electron beam evaporator and a standard Knudsen cell, respectively. The Nb:Se flux ratio was kept at 1:30, while evaporating the Se led to a pressure of ∼4 × 10–9 mbar (Se atmosphere). Samples were prepared using an evaporation time of 30 min to obtain a coverage of ∼1–2 ML. To minimize the presence of atomic defects, evaporation of Se was subsequently kept for additional 5 min. Atomic Force Microscopy at ambient conditions was routinely used to optimize the morphology of the NbSe2 layers. The samples used for AFM characterization were not further used for STM.

Lastly, in order to transfer the samples from our MBE to Madrid for HAS measurements, they were capped with a ∼ 10 nm film of Se. The capping layer was easily removed in the UHV-HAS system by annealing the sample at ∼300 °C. The bulk NbSe2(0001) surface measurements have been performed after ex-situ exfoliation of a bulk NbSe2(0001) sample.

Supplementary Material

nn4c16399_si_001.pdf (352.8KB, pdf)

Acknowledgments

This work has been partially supported by the Spanish Ministerio de Ciencia e Innovación under projects PID2019-109525RB-I00 and PID2023-147466OB-C21. D.F. acknowledges financial support from the Spanish Ministry of Economy and Competitiveness, through the “María de Maeztu” Programme for Units of Excellence in R&D (CEX2018-000805-M). M.M.U. acknowledges support by the ERC Starting grant LINKSPM (Grant 758558) and by grant PID2023-153277NB-I00 funded by the Spanish Ministry of Science, Innovation and Universities.

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsnano.4c16399.

  • Additional experimental details and complementary results. It provides further details on the two HAS setups used in our experiments. HAS in-plane and out-of-plane diffraction spectra of the Moiré pattern formed by NbSe2 are presented in Figure S1. Figure S2 shows in-plane and out-of-plane HAS angular distributions measured along ΓM from the NbSe2 (0001) surface (PDF)

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

The authors declare no competing financial interest.

References

  1. Naito M., Tanaka S.. Electrical Transport Properties in 2H-NbS2, -NbSe2, -TaS2 and -TaSe2 . J. Phys. Soc. Jpn. 1982;51:219–227. doi: 10.1143/JPSJ.51.219. [DOI] [Google Scholar]
  2. Foner S., McNiff E. J.. Upper critical fields of layered superconducting NbSe2 at low temperature. Phys. Lett. A. 1973;45:429–430. doi: 10.1016/0375-9601(73)90693-2. [DOI] [Google Scholar]
  3. Ugeda M. M., Bradley A. J., Zhang Y., Onishi S., Chen Y., Ruan W., Ojeda-Aristizabal C., Ryu H., Edmonds M. T., Tsai H. Z., Riss A., Mo S. K., Lee D., Zett A., Hussain Z., Shen Z. K., Crommie M. F.. Characterization of collective ground states in single-layer NbSe2 . Nat. Phys. 2016;12:92–97. doi: 10.1038/nphys3527. [DOI] [Google Scholar]
  4. Xi X., Zhao L., Wang Z., Berger H., Forró L., Shan J., Mak K. F.. Strongly enhanced charge-density-wave order in monolayer NbSe2 . Nat. Nanotechnol. 2015;10:765–769. doi: 10.1038/nnano.2015.143. [DOI] [PubMed] [Google Scholar]
  5. Semov W. I.. Work Function of Oxidized Metal Surfaces and Estimation of A1203 Film Band Structure Parameters. Phys. Status Solidi. 1969;32:K41. doi: 10.1002/pssb.19690320162. [DOI] [Google Scholar]
  6. Mammadov S., Ristein J., Krone J., Raidel Ch., Wanke M., Wiesmann V., Speck F., Seyller T.. Work function of graphene multilayers on SiC(0001) 2D Mater. 2017;4:015043. doi: 10.1088/2053-1583/4/1/015043. [DOI] [Google Scholar]
  7. Shimada T., Ohuchi F. S., Parkinson B. A.. Work Function and Photothreshold of Layered Metal Dichalcogenides. Jpn. J. Appl. Phys. 1994;33:2696. doi: 10.1143/JJAP.33.2696. [DOI] [Google Scholar]
  8. Benedek, G. ; Toennies, J. P. . Atomic-Scale Dynamics at Surfaces - Theory and Experimental Studies with Helium Atom Scattering; Springer-Nature, 2018. [Google Scholar]
  9. Li E., Hu J. X., Feng X., Zhou Z., An L., Law K. T., Wang N., Lin N.. Lattice reconstruction induced multiple ultra-flat bands in twisted bilayer WSe2 . Nat. Commun. 2021;12:5601. doi: 10.1038/s41467-021-25924-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Lisi S., Lu X., Benschop T., de Jong T. A., Stepanov P., Duran J. R., Margot F., Cucchi I., Cappelli E., Hunter A., Tamai A., Kandyba V., Giampietri A., Barinov A., Jobst J., Stalman V., Leeuwenhoek M., Watanabe K., Taniguchi T., Rademaker L., van der Molen S. J., Allan M. P., Efetov D. K., Baumberger F.. Observation of flat bands in twisted bilayer graphene. Nat. Phys. 2021;17:189–193. doi: 10.1038/s41567-020-01041-x. [DOI] [Google Scholar]
  11. Arora H. S., Polski R., Zhang Y., Thomson A., Choi Y., Kim H., Lin Z., Wilson I. Z., Xu X., Chu J. H., Watanabe K., Taniguchi T., Alicea J., Nadj-Perge S.. Superconductivity in metallic twisted bilayer graphene stabilized by WSe2 . Nature. 2020;583:379–381. doi: 10.1038/s41586-020-2473-8. [DOI] [PubMed] [Google Scholar]
  12. Manson J. R., Benedek G., Miret-Artés S.. Atom scattering as a probe of the surface electron-phonon interaction at conducting surfaces. Surf. Sci. Rep. 2022;77:100552. doi: 10.1016/j.surfrep.2022.100552. [DOI] [PubMed] [Google Scholar]
  13. Anemone G., Garnica M., Zappia M., Casado Aguilar P., Al Taleb A., Kuo C. N., Lue C. S., Politano A., Benedek G., Vázquez de Parga A. L., Miranda R., Farías D.. Experimental determination of surface thermal expansion and electron-phonon coupling constant of 1T-PtTe2 . 2D Mater. 2020;7:025007. doi: 10.1088/2053-1583/ab6268. [DOI] [Google Scholar]
  14. Anemone G., Casado Aguilar P., Garnica M., Calleja F., Al Taleb A., Kuo C. N., Lue C. S., Politano A., Vázquez de Parga A. L., Benedek G., Farías D., Miranda R.. Electron-phonon coupling in superconducting 1T-PdTe2 . npj 2D Mater. Appl. 2021;5:25. doi: 10.1038/s41699-021-00204-5. [DOI] [Google Scholar]
  15. Benedek G., Manson J. R., Miret-Artés S.. The Electron-Phonon Coupling Constant for Single-Layer Graphene on Metal Substrates Determined from He Atom Scattering. Phys. Chem. Chem. Phys. 2021;23:7575–7585. doi: 10.1039/D0CP04729E. [DOI] [PubMed] [Google Scholar]
  16. Ribeiro-Palau R., Zhang C., Watanabe K., Taniguchi T., Hone J., Dean C. R.. Twistable electronics with dynamically rotatable heterostructures. Science. 2018;361:690–693. doi: 10.1126/science.aat6981. [DOI] [Google Scholar]
  17. Minniti M., Díaz C., Fernández Cuñado J. L., Politano A., Maccariello D., Martín F., Farías D., Miranda R.. Helium, neon and argon diffraction from Ru(0001) J. Phys.: Condens. Matter. 2012;24:354002. doi: 10.1088/0953-8984/24/35/354002. [DOI] [PubMed] [Google Scholar]
  18. Zhu T., Litwin P. M., Rosul M. G., Jessup D., Akhanda M. S., Tonni F. F., Krylyuk S., Davydov A. V., Reinke P., McDonnell S. J., Zebarjadi M.. Transport properties of few-layer NbSe2: From electronic structure to thermoelectric properties. Mater. Today Phys. 2022;27:100789. doi: 10.1016/j.mtphys.2022.100789. [DOI] [Google Scholar]
  19. Yang G., Li L., Lee W. B., Ng M. C.. Structure of graphene and its disorders: a review. Sci. Technol. Adv. Mater. 2018;19(1):613–648. doi: 10.1080/14686996.2018.1494493. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. McCann E., Koshino M.. The electronic properties of bilayer graphene. Rep. Prog. Phys. 2013;76:056503. doi: 10.1088/0034-4885/76/5/056503. [DOI] [PubMed] [Google Scholar]
  21. Barredo D., Laurent G., Nieto P., Farías D., Miranda R.. High-resolution elastic and rotationally inelastic diffraction of D2 from NiAl(110) J. Chem. Phys. 2010;133:124702. doi: 10.1063/1.3479587. [DOI] [PubMed] [Google Scholar]
  22. Benedek G., Manson J. R., Miret-Artés S., Ruckhofer A., Ernst W. E., Tamtögl A., Toennies J. P.. Measuring the Electron-Phonon Interaction in Two-Dimensional Superconductors with He Atom Scattering. Condens. Matter. 2020;5:79. doi: 10.3390/condmat5040079. [DOI] [Google Scholar]
  23. Moncton D. E., Axe J. D., Di Salvo F. J.. Neutron scattering study of the charge-density wave transitions in 2H-TaSe 2 and 2H-NbSe 2. Phys. Rev. B. 1977;16:801. doi: 10.1103/PhysRevB.16.801. [DOI] [Google Scholar]
  24. Benedek G., Brusdeylins G., Heimlich C., Miglio L., Skofronick J., Toennies J. P.. Shifted Surface Phonon Anomaly in 2H-TaSe2(001) Phys. Rev. Lett. 1988;60:1037. doi: 10.1103/PhysRevLett.60.1037. [DOI] [PubMed] [Google Scholar]
  25. Al Taleb A., Anemone G., Farías D., Miranda R.. Resolving localized phonon modes on graphene/Ir(111) by inelastic atom scattering. Carbon. 2018;133:31–38. doi: 10.1016/j.carbon.2018.03.008. [DOI] [Google Scholar]
  26. Maccariello D., Campi D., Al Taleb A., Benedek G., Farías D., Bernasconi M., Miranda R.. Low-energy excitations of graphene on Ru(0001) Carbon. 2015;93:1–10. doi: 10.1016/j.carbon.2015.05.028. [DOI] [Google Scholar]
  27. Weber F., Hott R., Heid R., Bohnen K. P., Rosenkranz S., Castellan J. P., Osborn R., Said A. H., Leu B. M., Reznik D.. Optical phonons and the soft mode in 2H-NbSe2 . Phys. Rev. B. 2013;87:245111. doi: 10.1103/PhysRevB.87.245111. [DOI] [Google Scholar]
  28. Benedek G., Manson J. R., Miret-Artés S.. The Role of High-Energy Phonons in Electron-Phonon Interaction at Conducting Surfaces with Helium-Atom Scattering. Phys. Chem. Chem. Phys. 2022;24:23135–23141. doi: 10.1039/D2CP03501D. [DOI] [PubMed] [Google Scholar]
  29. Harper J. M. E., Geballe T. H., DiSalvo F. J.. Thermal properties of layered transition-metal dichalcogenides at charge-density-wave transitions. Phys. Rev. B. 1977;15:2943. doi: 10.1103/PhysRevB.15.2943. [DOI] [Google Scholar]
  30. Marezio M., Dernier P. D., Menth A., Hull Jr G. W.. The crystal structure of NbSe2 at 15 K. J. Solid State Chem. 1972;4:425–429. doi: 10.1016/0022-4596(72)90158-2. [DOI] [Google Scholar]
  31. Benedek G., Manson J. R., Miret-Artés S.. The Electron–Phonon Interaction of Low-Dimensional and Multi-Dimensional Materials from He Atom Scattering. Adv. Mater. 2020;32:2002072. doi: 10.1002/adma.202002072. [DOI] [PubMed] [Google Scholar]
  32. Anemone G., Al Taleb A., Benedek G., Castellanos-Gomez A., Farías D.. Electron-Phonon Coupling Constant of MoS2 . J. Phys. Chem. C. 2019;123:3682–3686. [Google Scholar]
  33. Valla T., Fedorov A. V., Johnson P. D., Glans P. A., McGuinness C., Smith K. E., Andrei E. Y., Berger H.. Quasiparticle Spectra, Charge-Density Waves, Superconduct-ivity, and Electron-Phonon Coupling in 2H–NbSe2 . Phys. Rev. Lett. 2004;92:086401. doi: 10.1103/PhysRevLett.92.086401. [DOI] [PubMed] [Google Scholar]
  34. Rahn D. J., Hellmann S., Kalläne M., Sohrt C., Kim T. K., Kipp L., Rossnagel K.. Gaps and kinks in the electronic structure of the superconductor 2H-NbSe2 from angle-resolved photoemission at 1 K. Phys. Rev. B. 2012;85:224532. doi: 10.1103/PhysRevB.85.224532. [DOI] [Google Scholar]
  35. Lian C. S., Si C., Wu J., Duan W.. First-principles study of Na-intercalated bilayer NbSe2: Suppressed charge-density wave and strain-enhanced superconductivity. Phys. Rev. B. 2017;96:235426. doi: 10.1103/PhysRevB.96.235426. [DOI] [Google Scholar]
  36. Zheng F., Feng J.. Electron-phonon coupling and the coexistence of super-conductivity and charge-density wave in monolayer NbSe2 . Phys. Rev. B. 2019;99:161119. doi: 10.1103/PhysRevB.99.161119. [DOI] [Google Scholar]
  37. Lian C. S., Si C., Duan W.. Unveiling Charge-Density Wave, Superconductivity, and Their Competitive Nature in Two-Dimensional NbSe2 . Nano Lett. 2018;18:2924. doi: 10.1021/acs.nanolett.8b00237. [DOI] [PubMed] [Google Scholar]
  38. McMillan W. L.. Transition Temperature of Strong-Coupled Superconductors. Phys. Rev. 1968;167:331. doi: 10.1103/PhysRev.167.331. [DOI] [Google Scholar]
  39. Dynes R. C.. McMillan’s equation and the Tc of superconductors. Solid State Commun. 1972;10:615–618. doi: 10.1016/0038-1098(72)90603-5. [DOI] [Google Scholar]
  40. Rossnagel K.. On the origin of charge-density waves in select layered transition-metal dichalcogenides. J. Phys.:Condens. Matter. 2011;23:213001. doi: 10.1088/0953-8984/23/21/213001. [DOI] [PubMed] [Google Scholar]
  41. Revolinsky E., Lautenschlager E. P., Armitage C. H.. Layer structure superconductor. Solid State Commun. 1963;1:59–61. doi: 10.1016/0038-1098(63)90358-2. [DOI] [Google Scholar]
  42. Cao Y., Mishchenko A., Yu G. L.. et al. Quality Heterostructures from Two-Dimensional Crystals Unstable in Air by Their Assembly in Inert Atmosphere. Nano Lett. 2015;15(8):4914–4921. doi: 10.1021/acs.nanolett.5b00648. [DOI] [PubMed] [Google Scholar]
  43. Wan W., Dreher P., Muñoz-Segovia D., Harsh R., Guo H., Martínez-Galera A. J., Guinea F., de Juan F., Ugeda M. M.. Observation of Superconducting Collective Modes fromCompeting Pairing Instabilities in Single-Layer NbSe2 . Adv. Mater. 2022;34:2206078. doi: 10.1002/adma.202206078. [DOI] [PubMed] [Google Scholar]
  44. Nakata Y., Sugawara K., Ichinokura S., Okada Y., Hitosugi T., Koretsune T., Ueno K., Hasegawa S., Takahashi T., Sato T.. Anisotropic band splitting in monolayer NbSe2: implications for superconductivity and charge density wave. npj 2D Mater. Appl. 2018;2:12. doi: 10.1038/s41699-018-0057-3. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

nn4c16399_si_001.pdf (352.8KB, pdf)

Articles from ACS Nano are provided here courtesy of American Chemical Society

RESOURCES